Paper The following article is Open access

Uniqueness and Lipschitz stability in electrical impedance tomography with finitely many electrodes

Published 3 January 2019 © 2019 IOP Publishing Ltd
, , Citation Bastian Harrach 2019 Inverse Problems 35 024005 DOI 10.1088/1361-6420/aaf6fc

0266-5611/35/2/024005

Abstract

For the linearized reconstruction problem in electrical impedance tomography with the complete electrode model, Lechleiter and Rieder (2008 Inverse Problems 24 065009) have shown that a piecewise polynomial conductivity on a fixed partition is uniquely determined if enough electrodes are being used. We extend their result to the full non-linear case and show that measurements on a sufficiently high number of electrodes uniquely determine a conductivity in any finite-dimensional subset of piecewise-analytic functions. We also prove Lipschitz stability, and derive analogue results for the continuum model, where finitely many measurements determine a finite-dimensional Galerkin projection of the Neumann-to-Dirichlet operator on a boundary part.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

We consider the inverse conductivity problem of determining the coefficient function $\sigma$ in the elliptic partial differential equation

Equation (1)

from knowledge of boundary measurements of $u$ . The problem arises in electrical impedance tomography (EIT), or electrical resistivity tomography, which is a novel technique to image the conductivity distribution $\sigma$ inside a subject $\Omega$ from electric voltage and current measurements on the subject's boundary $\partial \Omega$ , see [1, 11, 13, 20, 21, 26, 49, 50, 71, 73, 75, 77, 80, 87, 89], and the references therein for a broad overview on the developments in EIT.

To model the boundary measurements we consider the continuum model, where we measure the local Neumann-to-Dirichlet (NtD) operator (on a boundary part $\Sigma\subseteq \partial \Omega$ )

and the more realistic complete electrode model (CEM) with electrodes $E_1,\ldots,E_M\subseteq \partial \Omega$ all having the same contact impedance $z>0$ . In the CEM, we measure

where $u$ solves (1) with

The question whether full or local Neumann–Dirichlet-measurements uniquely determine the coefficient function $\sigma$ has become famous under the name Calderón problem [23, 24], and has been intensively studied in the mathematical literature due to its practical relevance for EIT and many other related inverse coefficient problems, see [7, 9, 25, 28, 34, 35, 54, 55, 58, 6163, 6668, 76, 84].

In this work we will study the question whether $\sigma$ can be uniquely and stably reconstructed from a finite number of electrode measurements. A natural discretization is to assume that $\sigma$ is piecewise constant (or piecewise polynomial) on a given resolution or partition of $\Omega$ , so that $\sigma$ will lie in an a priori known finite-dimensional subset $ \newcommand{\FF}{\mathcal{F}} \FF$ of piecewise-analytic functions. Moreover, it seems natural to assume that upper and lower bounds on the conductivity are a priori known, i.e.

Our main result for the continuum model is that a (sufficiently high dimensional) finite-dimensional Galerkin projection ${P_{G_N}} \Lambda(\sigma) {P_{G_N}}$ already uniquely determines $\sigma$ and that Lipschitz stability holds

see theorem 2.4.

Under the additional assumption that $\sigma$ is an a priori known smooth function close to the boundary, we then turn to the CEM. We show that a (sufficiently large) finite number of electrodes $M$ suffices to uniquely determine $\sigma$ with Lipschitz stability

see theorem 3.1. This shows that the discretized EIT problem is uniquely and stably solvable if enough electrodes are being used, which may be relevant for practical implementations of EIT reconstruction algorithms.

Note that our results are non-constructive, we do not have a practically useful estimate of the Lipschitz constant or the required number of electrodes yet. Also note, that the necessary number of electrodes and the stability constant $c>0$ depend on the ansatz set $ \newcommand{\FF}{\mathcal{F}} \FF_{[a,b]}$ . Due to the intrinsic ill-posedness of the non-discretized EIT problem, we can naturally expect that a larger set $ \newcommand{\FF}{\mathcal{F}} \FF_{[a,b]}$ will lead to worse stability constants and a higher required number of electrodes, with $c\to 0$ and $M\to \infty$ when $ \newcommand{\FF}{\mathcal{F}} \mathrm{dim}({\rm span}~ \FF)\to \infty$ .

Let us give some more references on related results and the origins of our approach. A recent preprint of Alberti and Santacesaria [2] uses complex geometrical optics solutions to show that (in the continuuum model) there exists a finite number of boundary voltages, so that the knowledge of the corresponding boundary currents uniquely determines the conductivity $\sigma$ and that Lipschitz stability holds. Their result holds in dimension $d\geqslant 3$ with measurements on the full boundary $\partial \Omega$ , $\sigma$ is assumed to be identically one close to $\partial \Omega$ , bounded by a priori known constants, and $\frac{\Delta {\sqrt{\sigma}}}{{\sqrt{\sigma}}}$ has to belong to an a priori known finite-dimensional subspace of $L^\infty$ . Our result in this work requires less restrictive assumptions as we can treat any dimension $d\geqslant 2$ , partial boundary data, and the CEM. But, on the other hand, we require the assumption of piecewise-analyticity which is more restrictive than the assumptions in [2].

For the linearized EIT problem (both, in the continuum model, and with the CEM), Lechleiter and Rieder [70] have shown that a piecewise polynomial conductivity on a fixed partition is uniquely determined if enough electrodes are being used. The main tool in [70] is the theory of localized potentials devoloped by the author [32] and the convergence of CEM-solutions to solutions of the continuum model shown by Hyvönen, Lechleiter and Hakula [51, 69]. Our result uses similar tools and first treats the non-linear EIT problem with the continuum model using localized potentials [32, 46] and monotonicity estimates between the non-linearized and the linearized problem from Ikehata, Kang, Seo and Sheen [53, 59]. Then we extend the results to the CEM using recent results on the approximation of the continuum model by the CEM from Hyvönen, Garde and Staboulis [30, 52].

The idea of using monotonicity estimates and localized potentials techniques has lead to a number of results for inverse coefficient problems [8, 12, 22, 33, 36, 37, 39, 40, 4446, 48], and several recent works build practical reconstruction methods on monotonicity properties [2931, 38, 42, 43, 47, 72, 83, 85, 86, 88, 92]. Together with the recent preprints [41, 79], the present work shows that this idea can also be used to obtain Lipschitz stability estimates, which are usually derived from technically more challenging approaches involving Carleman estimates or quantitative unique continuation, see [36, 10, 1419, 27, 56, 57, 60, 64, 65, 74, 78, 81, 90, 91].

The work is organized as follows. In section 2 we treat the continuum model, and show that the NtD operator or a (sufficiently high dimensional) finite-dimensional Galerkin projection uniquely determine the conductivity with Lipschitz stability. We formulate our main results for the continuum model in theorems 2.3 and 2.4 in section 2.1, summarize some known results from the literature in section 2.2, and the prove the theorems in section 2.3. In section 3 we then treat the CEM. Again we first formulate a uniqueness and Lipschitz stability result in theorem 3.1 in section 3.1, then summarize known results from the literature in section 3.2, and finally prove the theorem in section 3.3.

2. Uniqueness and Lipschitz stability from continuous data

2.1. Setting and main results

Let $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \Omega\subset\R^d$ , $d\geqslant 2$ be a bounded domain with smooth boundary $\partial\Omega$ and outer normal vector $\nu$ . $L_+^\infty(\Omega)$ denotes the subspace of $L^\infty(\Omega)$ -functions with positive essential infima. $ \newcommand{\Hd}{H_\diamond} \Hd^1(\Omega)$ and $ \newcommand{\Ld}{L_\diamond} \Ld^2(\partial \Omega)$ denote the spaces of $H^1$ - and $L^2$ -functions with vanishing integral mean on $\partial \Omega$ .

For $\sigma\in L_+^\infty(\Omega)$ , and a relatively open boundary part $\Sigma\subseteq \partial \Omega$ , the local NtD operator $\Lambda(\sigma)$ is defined by

where $u^g_\sigma\in H^1_\diamond(\Omega)$ is the unique solution of

Equation (2)

This is equivalent to the variational formulation that $u^g_\sigma\in H^1_\diamond(\Omega)$ solves

Equation (3)

It is well known and easily shown that $\Lambda(\sigma)$ is compact and self-adjoint.

We will consider conductivities that are a priori known to belong to a finite dimensional set of piecewise-analytic functions and that are bounded from above and below by a priori known constants. To that end, we first define piecewise-analyticity as in [46, definition 2.1]:

Definition 2.1. 

  • (a)  
    A subset $\Gamma\subseteq \partial O$ of the boundary of an open set $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} O\subseteq \R^n$ is called a smooth boundary piece if it is a $C^\infty$ -surface and $O$ lies on one side of it, i.e. if for each $z\in \Gamma$ there exists a ball $ \newcommand{\e}{{\rm e}} B_\epsilon(z)$ and a function $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \gamma\in C^\infty(\R^{n-1},\R)$ such that upon relabeling and reorienting
  • (b)  
    $O$ is said to have smooth boundary if $\partial O$ is a union of smooth boundary pieces. $O$ is said to have piecewise smooth boundary if $\partial O$ is a countable union of the closures of smooth boundary pieces.
  • (c)  
    A function $\kappa \in L^\infty(\Omega)$ is called piecewise analytic if there exist finitely many pairwise disjoint subdomains $O_1,\ldots,O_M\subset \Omega$ with piecewise smooth boundaries, such that $\overline{\Omega}= \overline{O_1\cup \ldots \cup O_M}$ , and $\kappa|_{O_m}$ has an extension which is (real-)analytic in a neighborhood of $\overline{O_m}$ , $m=1,\ldots,M$ .

Note that (to the knowledge of the author), it is not clear whether the sum of two piecewise-analytic functions is always piecewise-analytic, i.e. whether the set of piecewise-analytic functions is a vector space. However, this can be guaranteed with a slightly stronger definition of piecewise analyticity (see [67, lemma 1]). Moreover, finite-dimensional vector spaces of piecewise-analytic functions (or subsets thereof) naturally arise as parameter spaces for the inverse conductivity problem, e.g. when we fix a partition of the imaging domain $\Omega$ into a finite number of subdomains (e.g. triangles, pixels, or voxels) and the conductivity is assumed to be a polynomial of fixed maximal order on each of these subdomains. Therefore, we make the following definition:

Definition 2.2. A set $ \newcommand{\FF}{\mathcal{F}} \FF\subseteq L^\infty(\Omega)$ is called a finite-dimensional subset of piecewise-analytic functions if its linear span

contains only piecewise-analytic functions and $ \newcommand{\FF}{\mathcal{F}} \mathrm{dim}({\rm span}~ \FF)<\infty$ .

Given a finite-dimensional subset $ \newcommand{\FF}{\mathcal{F}} \FF$ of piecewise analytic functions and two numbers $b>a>0$ , we denote the set

Throughout this paper, the domain $\Omega$ , the finite-dimensional subset $ \newcommand{\FF}{\mathcal{F}} \FF$ and the bounds $b>a>0$ are fixed, and the constants in the Lipschitz stability results will depend on them.

Our first result shows Lipschitz stability for the inverse conductivity problem in $ \newcommand{\FF}{\mathcal{F}} \FF_{[a,b]}$ when the complete infinite-dimensional NtD-operator is measured.

Theorem 2.3. There exists $c>0$ such that

Proof. Theorem 2.3 will be proven in section 2.3. □

We then turn to the question whether $ \newcommand{\FF}{\mathcal{F}} \sigma\in \FF_{[a,b]}$ is already uniquely determined by finitely many boundary measurements in the continuum model. For a (finite- or infinite-dimensional) subspace $ \newcommand{\Ld}{L_\diamond} G\subseteq \Ld^2(\Sigma)$ we denote by

the orthogonal projection operator on $G$ with respect to the $L^2$ -scalar product

Equation (4)

Clearly, $P_G=P_G^*$ , and if $G$ is finite dimensional with a basis $ \newcommand{\re}{{\rm Re}} \newcommand{\psan}{\mathrm{span}\,} G=\psan(g_1,\ldots,g_n)$ then measurements of

determine the Galerkin projection of the NtD operator $P_{G} \Lambda(\sigma) {P_G}$ , so that this can be regarded as a model for finitely many voltage/current measurements in the continuum model.

Our next result shows that this uniquely determines $ \newcommand{\FF}{\mathcal{F}} \sigma\in \FF_{[a,b]}$ (with Lipschitz stability) if the space $G$ is large enough.

Theorem 2.4. For each sequence of subspaces

there exists $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} N\in \N$ , and $c>0$ such that

for all $ \newcommand{\FF}{\mathcal{F}} \sigma_1,\sigma_2\in \FF_{[a,b]}$ , and all $n\geqslant N$ .

In particular, this implies that for all $ \newcommand{\FF}{\mathcal{F}} \sigma_1,\sigma_2\in \FF_{[a,b]}$ and all $n\geqslant N$

Proof. Theorem 2.4 will be proven in section 2.3. □

2.2. Differentiability, monotonicity and localized potentials

In this subsection, we summarize some known results from the literature, that we will use to prove theorem 2.3 and 2.4. As defined in (4), $\langle \cdot,\cdot \rangle$ always denotes the $L^2(\Sigma)$ -scalar product, and $u^g_\sigma\in H^1_\diamond(\Omega)$ denotes the solution of (2) with conductivity $\sigma\in L^\infty_+(\Omega)$ and Neumann data $ \newcommand{\Ld}{L_\diamond} g\in \Ld^2(\Sigma)$ in the following.

Our first tool is that the NtD operator is continuously Fréchet differentiable with respect to the conductivity.

Lemma 2.5. 

  • (a)  
    The mapping
    is Fréchet differentiable. Its derivative is given by
    Equation (5)
    where $ \newcommand{\Hd}{H_\diamond} v\in \Hd^1(\Omega)$ solves
  • (b)  
    For all $\sigma\in L^\infty_+(\Omega)$ and $\kappa\in L^\infty(\Omega)$ the operator $ \newcommand{\Ld}{L_\diamond} \newcommand{\LL}{\mathcal{L}} \Lambda'(\sigma)\kappa\in \LL(\Ld^2(\Sigma))$ is self-adjoint and compact, and it fulfills
    for all $ \newcommand{\Ld}{L_\diamond} g,h\in \Ld^2(\Sigma)$ .
  • (c)  
    The mapping
    is continuous.

Proof. This follows from the variational formulation of the conductivity equation (3), see e.g. [70, section 2] or [30, appendix B]. □

Our next tool is a monotonicity relation between the NtD-operator and its derivative that goes back to Ikehata, Kang, Seo, and Sheen [53, 59], and has been used in several other works, see the list of works on monotonicity-based methods cited in the introduction.

Lemma 2.6. For all $\sigma_1,\sigma_2\in L^\infty_+(\Omega)$ and $g\in L^2_\diamond(\Sigma)$ , it holds that

Equation (6)

Proof. See, e.g. [45, lemma 2.1]. □

The energy terms $|\nabla u_\sigma^g|^2$ in the monotonicity estimate can be controlled using the technique of localized potentials [32]. Roughly speaking, the energy $|\nabla u_\sigma^g|^2$ can be made arbitrarily large in a subset $D_1\subseteq \Omega$ without making it large in another subset $D_2\subseteq \Omega$ whenever $D_1$ can be reached from the boundary without passing $D_2$ .

To formulate this rigorously, we adopt the notation from [46, definitions 2.2 and 2.3] and denote by $ \newcommand{\inn}{\mathrm{int}\,} \inn D$ the topological interior of a subset $D\subseteq \overline\Omega$ , and by $ \newcommand{\out}{\mathrm{out}_{\Sigma}\,} \out D$ its outer hull, i.e.

With this notation, we have the following localized potentials result:

Lemma 2.7. Let $\sigma\in L^\infty_+(\Omega)$ be piecewise analytic and let $D_1,D_2\subseteq\overline\Omega$ be two measurable sets with

Then there exists a sequence of currents $(g_m)_{m\in\mathbb {N}}\subset L^2_\diamond(\Sigma)$ such that the corresponding solutions $ \newcommand{\Hd}{H_\diamond} (u_\sigma^{g_m})_{m\in \mathbb {N}}\subset \Hd^1(\Omega)$ fulfill

Proof. ([46], theorem 3.6 and section 4.3).

proof 

We will also need the following definiteness property of piecewise-analytic functions from [46]:

Lemma 2.8. Let $ \newcommand{\e}{{\rm e}} 0\not\equiv \kappa\in L^\infty(\Omega)$ be piecewise-analytic. Then there exist two sets

with

and either

  • (i)  
    $\kappa|_{\Omega\setminus D_2}\geqslant 0$ and $\kappa|_{D_1}\in L_+^\infty(D_1)$ , or
  • (ii)  
    $\kappa|_{\Omega\setminus D_2}\leqslant 0$ and $-\kappa|_{D_1}\in L_+^\infty(D_1)$ .
Proof. ([46], theorem A.1, corollary A.2, and section 4.3).

proof 

2.3. Proof of theorems 2.3 and 2.4

We can now prove theorems 2.3 and 2.4. For the sake of brevity, we write $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \norm{\cdot}$ for $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \newcommand{\Ld}{L_\diamond} \newcommand{\LL}{\mathcal{L}} \norm{\cdot}_{\LL(\Ld^2(\Sigma))}$ , $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \norm{\cdot}_{L^\infty(\Omega)}$ and $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \newcommand{\Ld}{L_\diamond} \norm{\cdot}_{\Ld^2(\Sigma)}$ throughout this subsection.

We follow the approach in [41] and first use the monotonicity relation in lemma 2.6 to bound the difference of the non-linear NtD operators by an expression containing their linearized counterparts.

Lemma 2.9. For all $ \newcommand{\FF}{\mathcal{F}} \sigma_1,\sigma_2\in \FF_{[a,b]}$ with $ \newcommand{\e}{{\rm e}} \sigma_1\not\equiv \sigma_2$ ,

where $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \newcommand{\Ld}{L_\diamond} f:\ L^\infty_+(\Omega)\times L^\infty_+(\Omega)\times L^\infty(\Omega)\times \Ld^2(\Sigma)\to \R$ is defined by

and $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} \newcommand{\re}{{\rm Re}} \newcommand{\psan}{\mathrm{span}\,} \KK:=\{\kappa\in \psan\FF:\ \norm{\kappa}=1\}$ .

Proof. The NtD-operators are self-adjoint so that for all $\sigma_1,\sigma_2\in L^\infty(\Omega)$

Using the monotonicity inequality in lemma 2.6 also with interchanged roles of $\sigma_1$ and $\sigma_2$ , we obtain that for all $\sigma_1,\sigma_2\in L^\infty_+(\Omega)$ , $ \newcommand{\e}{{\rm e}} \sigma_1\not\equiv \sigma_2$ , and all $ \newcommand{\Ld}{L_\diamond} g\in \Ld^2(\partial \Omega)$

Hence,

Now we use a compactness argument to show that the expression in the lower bound in lemma 2.9 attains its minimum.

Lemma 2.10. There exists $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} (\hat \tau_1,\hat \tau_2,\hat \kappa) \in \FF_{[a,b]}\times \FF_{[a,b]}\times \KK$ so that

Proof. Since $f$ is continuous by lemma 2.5, the function

is lower semicontinuous and thus attains its minimum over the compact set $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} \FF_{[a,b]}\times $ $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} \FF_{[a,b]}\times \KK$ . □

It remains to show that the minimum attained in lemma 2.10 must be positive. To show that we use the localized potentials from lemma 2.7.

Lemma 2.11. Let $ \newcommand{\e}{{\rm e}} 0\not\equiv \kappa\in L^\infty(\Omega)$ be piecewise-analytic. Then at least one of the following two properties holds true:

  • (i)  
    For all piecewise analytic $\sigma\in L^\infty_+(\Omega)$ there exists $ \newcommand{\Ld}{L_\diamond} g\in \Ld^2(\partial \Omega)$ with
  • (ii)  
    For all piecewise analytic $\sigma\in L^\infty_+(\Omega)$ there exists $ \newcommand{\Ld}{L_\diamond} g\in \Ld^2(\partial \Omega)$ with

Hence, a fortiori,

Proof. Using the definiteness property of piecewise analytic functions from lemma 2.8 we obtain two sets $ \newcommand{\inn}{\mathrm{int}\,} D_1=\inn D_1$ and $ \newcommand{\out}{\mathrm{out}_{\Sigma}\,} D_2=\out D_2$ with

and either

  • (i)  
    $\kappa|_{\Omega\setminus D_2}\geqslant 0$ and $\kappa|_{D_1}\in L_+^\infty(D_1)$ , or
  • (ii)  
    $\kappa|_{\Omega\setminus D_2}\leqslant 0$ and $-\kappa|_{D_1}\in L_+^\infty(D_1)$ .

Let $(g_m)_{m\in\mathbb {N}}\subset L^2_\diamond(\Sigma)$ be the localized potentials sequence from lemma 2.7. Then, in case (i), we obtain

so that $ \newcommand{\e}{{\rm e}} \newcommand{\dx}[1][x]{{\,{{\rm d}} #1}} \int_\Omega \kappa |\nabla u_\sigma^{g_m}|^2 \dx>0$ for sufficiently large $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} m\in \N$ .

In case (ii) we obtain

so that $ \newcommand{\e}{{\rm e}} \newcommand{\dx}[1][x]{{\,{{\rm d}} #1}} \int_\Omega \kappa |\nabla u_\sigma^{g_m}|^2 \dx<0$ for sufficiently large $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} m\in \N$ . □

Remark 2.12. It is known (see e.g. [45, corollary 3.5(b)]) that for all piecewise analytic $\sigma$ , the Fréchet derivative $\Lambda'(\sigma)$ is injective on the set of piecewise analytic functions, i.e. $\Lambda'(\sigma)\kappa\neq 0$ for all piecewise analytic $ \newcommand{\e}{{\rm e}} 0\not\equiv \kappa\in L^\infty(\Omega)$ .

Since $\Lambda'(\sigma)\kappa$ is a compact self-adjoint operator, this means that $\Lambda'(\sigma)\kappa$ must possess either a positive or a negative eigenvalue. Lemma 2.11 can be interpreted in the sense, that for each $ \newcommand{\e}{{\rm e}} \kappa\not\equiv 0$ this property is sign-uniform in $\sigma$ , i.e. for each $ \newcommand{\e}{{\rm e}} \kappa\not\equiv 0$ , the operator $\Lambda'(\sigma)\kappa$ either possesses a positive eigenvalue for all $\sigma$ , or it possesses a negative eigenvalue for all $\sigma$ (or both properties are fulfilled).

With these preparations we can now show the theorems 2.3 and 2.4.

Proof of theorem 2.3. The assertion follows from lemmas 2.92.11 with

Proof of theorem 2.4. Using that

we obtain as in lemmas 2.9 and 2.10 that for all $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} n\in \N$ , there exists $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} (\hat \tau_1^{(n)},\hat \tau_2^{(n)},\hat \kappa^{(n)})$ $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} \in \FF_{[a,b]}\times \FF_{[a,b]}\times \KK$ so that

Equation (7)

The right hand side of (7) is monotonically increasing in $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} n\in \N$ since the spaces $G_n$ are nested. Hence, the assertion of theorem 2.4 follows, if we can prove that there exists $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} n\in \N$ with

Equation (8)

We argue by contradiction and assume that this is not the case. Then there exists a sequence $ \newcommand{\KK}{\mathcal{K}} (\tau_1^{(n)},\tau_2^{(n)},\kappa^{(n)})\in {\mathcal F}_{[a,b]}\times {\mathcal F}_{[a,b]}\times \KK$ with

which also implies

After passing to a subsequence if necessary, we can assume by compactness that the sequence $(\tau_1^{(n)},\tau_2^{(n)},\kappa^{(n)})$ converges to some element

Since, for all $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} m\in \N$ , the function

is lower semicontinuous, it follows that

But, by continuity, this would imply

which contradicts lemma 2.11. This shows that (8) must be true for sufficiently large $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} n\in \N$ and thus theorem 2.4 is proven. □

3. Uniqueness and Lipschitz stability from electrode measurements

3.1. Setting and main results

Now we consider the CEM. As before let $\sigma\in L^\infty_+(\Omega)$ denote the conductivity distribution in a smoothly bounded domain $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \Omega\subseteq \R^d$ , $d\geqslant 2$ . We assume that $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} M\in \N$ open, connected, mutually disjoint electrodes $E_m\subseteq \partial \Omega$ , $m=1,\ldots,M$ , are attached to the boundary of $\Omega$ with all electrodes having the same contact impedance $z>0$ . When a current with strength $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} J_m\in \R$ is driven through the $m$ -th electrode (with $\sum_{m=1}^M J_m=0$ ), the resulting electrical potential $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} (u,U)\in H^1(\Omega)\times \R^M$ solves the following equations:

Equation (9)

Equation (10)

Equation (11)

Equation (12)

where $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} U=(U_1,\ldots,U_M)\in \R^M$ is a vector containing the electric potentials on the electrodes $E_1,\ldots,E_M$ .

It can be shown that (9)–(12) possess a solution $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} (u,U)\in H^1(\Omega)\times \R^m$ and that the solution is unique under the additional gauge (or ground level) condition $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} U\in \Rd^M$ , where $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} \Rd^M$ is the subspace of vectors in $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \R^m$ with zero mean, see e.g. [82]. We can thus define the $M$ -electrode current-to-potential operator

where $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} (u,U)\in H^1(\Omega)\times \Rd^m$ solves (9)–(12). Note also that (9)–(12) are equivalent to the variational formulation that $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} (u,U)\in H^1(\Omega)\times \Rd^m$ solves

Equation (13)

for all $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} (w,W)\in H^1(\Omega)\times \Rd^m$ , see again, [82].

As in the previous section, we will consider conductivities that belong to a finite dimensional subset of piecewise-analytic functions. Additionally, in order to use results from [30] on the approximation properties of the CEM, we assume that the background conductivity is an a priori known smooth function in a fixed neighborhood $O$ of the boundary $\partial \Omega$ , i.e. we assume that $ \newcommand{\FF}{\mathcal{F}} \FF$ is a finite dimensional subset of piecewise-analytic functions, so that there exists $\sigma_0\in C^\infty({\overline O})$ with $\sigma|_{{O}}=\sigma_0|_{{O}}$ for all $ \newcommand{\FF}{\mathcal{F}} \sigma\in \FF$ . Together with the assumption of a priori known bounds, we assume (for $b>a>0$ )

We will show that $R_M(\sigma)$ uniquely determines $ \newcommand{\FF}{\mathcal{F}} \sigma\in\FF_{[a,b]}$ (with Lipschitz stability) if, roughly speaking, enough electrodes are being used. To make this statement precise, assume that the number of electrodes is increased so that the electrode configurations fulfill the Hyvönen criteria [30, 52]:

  • (H1)  
    For all electrode configurations
    there exist open, connected, and mutually disjoint sets (called virtual extended electrodes) $\tilde E_m^{(M)}$ , $m=1,\ldots,M$ with
    so that
  • (H2)  
    The operators
    fulfill that
    for all $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} M\in \N$ and all $f\in H^1(\partial \Omega)$ .

The first criterion implies the natural assumption that the electrode sizes shrink to zero, but always cover a certain fraction of the boundary. The somewhat technical second criterion can be interpreted as a Poincaré-type inequality that is fulfilled for regular enough electrode shapes, see [30, 52, 69]. Together these criteria guarantee that the electrode measurement approximate all possible continuous measurements in a suitable sense.

Now we can state our main result:

Theorem 3.1. There exists $ \newcommand{\e}{{\rm e}} \newcommand{\N}{{\mathbb {N}}} N\in \N$ and $c>0$ such that for all $M\geqslant N$

In particular, this implies that for all $ \newcommand{\FF}{\mathcal{F}} \sigma_1,\sigma_2\in \FF_{[a,b]}$ and $M\geqslant N$

Theorem 3.1 will be proven in section 3.3.

3.2. Differentiability, monotonicity, and approximation of linearized measurements

The electrode measurements $R_M(\sigma)$ fulfill analogue differentiability and monotonicity properties as the NtD-Operators. In the following $\langle \cdot,\cdot \rangle_M$ denotes the Euclidian scalar product in $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \R^M$ . For a vector $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} J=(J_1,\ldots,J_M)\in \Rd^M$ , we denote by $u^{(J)}_\sigma\in H^1_\diamond(\Omega)$ the solution of the CEM equations (9)–(12) with conductivity $\sigma\in L^\infty_+(\Omega)$ and electrode currents $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} J_1,\ldots,J_m\in \R$ .

Lemma 3.2. 

  • (a)  
    The mapping
    is Fréchet differentiable. Its derivative is given by
    where $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} (v,V)\in H^1(\Omega)\times \Rd^m$ solves
    Equation (14)
    for all $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} (w,W)\in H^1(\Omega)\times \Rd^m$ .
  • (b)  
    For all $\sigma\in L^\infty_+(\Omega)$ and $\kappa\in L^\infty(\Omega)$ the operator $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} \newcommand{\LL}{\mathcal{L}} R_M'(\sigma)\kappa\in \LL(\Rd^M)$ is self-adjoint, and it fulfills
    for all $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} I,J\in \Rd^M$ .
  • (c)  
    The mapping
    is continuous.

Proof. This follows from the variational formulation of the CEM (13), see e.g. [70, section 2] or [30, appendix B]. □

Lemma 3.3. For all $\sigma_1,\sigma_2\in L^\infty_+(\Omega)$ and $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} J\in \Rd^M$ , it holds that

Proof. ([47], theorem 2).

proof 

We will also require the following result from Garde and Staboulis [30] that the linearized CEM measurements approximate the linearized NtD operator.

Lemma 3.4. Under the Hyvönen assumptions (H1) and (H2), there exists $C>0$ such that for all $ \newcommand{\FF}{\mathcal{F}} \sigma\in \FF_{[a,b]}$ and $ \newcommand{\FF}{\mathcal{F}} \newcommand{\re}{{\rm Re}} \newcommand{\psan}{\mathrm{span}\,} \kappa\in \psan \FF$

where

Moreover, for all $ \newcommand{\Ld}{L_\diamond} g\in \Ld^2(\partial \Omega)$

Proof. ([30], theorem 3, proposition 4).

proof 

3.3. Proof of theorem 3.1

Again, for the sake of brevity, we omit norm subscripts when the choice of the norm is clear from the context. As in section 2.3 we obtain from the monotonicity result lemma 3.3 that

where $ \newcommand{\R}{{\mathbb {R}}} \newcommand{\e}{{\rm e}} \newcommand{\Rd}{{\mathbb {R}_\diamond}} f_M:\ L^\infty_+(\Omega)\times L^\infty_+(\Omega)\times L^\infty(\Omega)\times \Rd^M\to \R$ is defined by

and $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} \newcommand{\re}{{\rm Re}} \newcommand{\psan}{\mathrm{span}\,} \KK:=\{\kappa\in \psan \FF:\ \norm{\kappa}=1\}$ .

We compare this with $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} \newcommand{\Ld}{L_\diamond} f:\ L^\infty_+(\Omega)\times L^\infty_+(\Omega)\times L^\infty(\Omega)\times \Ld^2(\partial \Omega)\to \R$ ,

from the continuum model (see lemma 2.9) with $\Sigma=\partial \Omega$ .

For all real numbers $ \newcommand{\e}{{\rm e}} \newcommand{\R}{{\mathbb {R}}} r_1,r_2,s_1,s_2\in \R$ one has that

Hence, we obtain with lemma 3.4 that for all $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} (\tau_1,\tau_2,\kappa) \in \FF_{[a,b]}\times \FF_{[a,b]}\times \KK$ and $ \newcommand{\Ld}{L_\diamond} g\in \Ld^2(\partial \Omega)$ with $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \norm{g}=1$

where $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \norm{\kappa}=1$ and $ \newcommand{\norm}[1]{\hspace{0.2ex} \|#1\| \hspace{0.2ex}} \max\{\norm{\tau_1},\norm{\tau_2}\}\leqslant b$ follows from the definition of $ \newcommand{\KK}{\mathcal{K}} \KK$ and $ \newcommand{\FF}{\mathcal{F}} \FF_{[a,b]}$ .

For all $g\in {L^2(\partial \Omega)}$ we have that

so that we obtain for all $ \newcommand{\FF}{\mathcal{F}} \newcommand{\KK}{\mathcal{K}} (\tau_1,\tau_2,\kappa) \in \FF_{[a,b]}\times \FF_{[a,b]}\times \KK$

Since the first summand is positive by lemma 2.11, it follows that for sufficiently large numbers of electrodes $M$

With the same lower semicontinuity and compactness argument as in the continuum model, this yields

so that the assertion is proven. □

Acknowledgments

This work is devoted to the memory of Professor Armin Lechleiter who will be deeply missed as a scientist, and as a friend.

Please wait… references are loading.